Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jan 1.
Published in final edited form as: Stat Sin. 2009;19(3):1077–1093.

Spatial Linear Mixed Models with Covariate Measurement Errors

Yi Li 1,3, Haicheng Tang 2, Xihong Lin 3
PMCID: PMC2695401  NIHMSID: NIHMS94321  PMID: 20046975

SUMMARY

Spatial data with covariate measurement errors have been commonly observed in public health studies. Existing work mainly concentrates on parameter estimation using Gibbs sampling, and no work has been conducted to understand and quantify the theoretical impact of ignoring measurement error on spatial data analysis in the form of the asymptotic biases in regression coefficients and variance components when measurement error is ignored. Plausible implementations, from frequentist perspectives, of maximum likelihood estimation in spatial covariate measurement error models are also elusive. In this paper, we propose a new class of linear mixed models for spatial data in the presence of covariate measurement errors. We show that the naive estimators of the regression coefficients are attenuated while the naive estimators of the variance components are inflated, if measurement error is ignored. We further develop a structural modeling approach to obtaining the maximum likelihood estimator by accounting for the measurement error. We study the large sample properties of the proposed maximum likelihood estimator, and propose an EM algorithm to draw inference. All the asymptotic properties are shown under the increasing-domain asymptotic framework. We illustrate the method by analyzing the Scottish lip cancer data, and evaluate its performance through a simulation study, all of which elucidate the importance of adjusting for covariate measurement errors.

Keywords: Measurement error, Spatial data, Structural modeling, Variance components, Asymptotic bias, Consistency and asymptotic normality, Increasing domain asymptotics, EM algorithm

1 Introduction

Spatial data are common in ecology, environmental health and epidemiology, where sampling units are geographical areas or spatially located individuals (Cressie (1993)). Analysis of spatial data is challenged by the spatial correlation among the observations. Mixed effects models provide a convenient framework to model the spatial correlation using random effects that are assumed to follow some spatial correlation structure, such as the conditional autoregressive (CAR) structure (Yasui and Lele (1997); Waller, et al. (1997)) or the Matèrn correlation structure (Stein (1999)). Asymptotic theory for spatial linear mixed models was established by Mardia and Marshall (1984).

Spatial data are susceptible to measurement errors in covariates. For example in ecological studies, covariates are often collected from a small survey sample in each area and sample averages are used as surrogates for the true population aggregated values, such as the percentage of smokers in a county (Xia and Carlin, 1998). Measurement errors can be substantial when the areas are small, especially in nutritional ecological studies (Prentice and Sheppard (1995)), where additional measurement errors arise due to inaccuracy in measuring nutrition intakes, such as fat intake, using conventional instruments and using 24 hour food recall. In environmental health studies, the air pollution level, e.g., PM10 or ozone, in an area is difficult to measure and is often approximated by using the distance from a polluted site or by using the measures at a few monitoring sites (Carroll, et al., 1997).

There is a vast literature on modeling measurement error for independent data. For an overview, see Carroll, Stefanski and Ruppert (1995). Several authors have considered measurement error in covariate in generalized linear mixed models for clustered data, such as longitudinal data (Wang and Davidian (1996); Wang, et al. (1998)). However, only limited work has been done in modeling measurement error in covariates for spatial data. Bernardinelli, et al. (1997) and Xia and Carlin (1998) accounted for measurement error in covariate using hierarchical models in disease mapping. These authors mainly concentrated on parameter estimation using Gibbs sampling. Little is understood about the theoretical effect of measurement error on the asymptotic biases in regression coefficients and variance components when measurement error is ignored. To our knowledge, our work is the first attempt to understand the theoretical properties of maximum likelihood estimation in spatial measurement error mixed effects models.

We first study the asymptotic bias in the naive estimator when measurement error is ignored. Our results show that ignoring measurement error results in attenuated regression coefficients and inflated variance components. We then proceed by applying the structural modeling approach to make valid maximum likelihood inference by accounting for measurement error. An EM algorithm is proposed to compute the maximum likelihood estimator. The proposed methods are illustrated through an application to the Scottish lip cancer data and their performance is evaluated through a simulation study.

2 The Spatial Linear Mixed Measurement Error Model

Suppose that the data are obtained from n geographical areas with continuous outcome variable Yi, unobserved true covariate Xi (assumed to be a scalar), observed Xi-related covariate Wi, and other accurately observed covariates Zi at the ith area (i = 1,…, n). Conditional on the site-specific random effects bi that model the spatial correlation, the spatial linear mixed model of Y given X and Z can be written as

Yi=β0+Xiβx+ZiTβz+bi+εi, (1)

where the random effect vector (b1,…, bn) is N{0, V(θ)} and θ is a vector of variance components, the residuals εi are N(0, σε2), and bi and εi are independent to each other and are independent of the covariates X and Z.

The covariance matrix V(θ) models the spatial correlation and admits many choices. For instance, we might parameterize the (i, j)th component of V(θ) as Vij(θ) = θρ(||sisj||), where correlation function ρ(.) is an isotropic correlation function that decays as the Euclidean distance dij = ||sisj|| between two individuals increases. A widely adopted choice for this correlation function is the Matèrn function 12ν1Γ(ν)(2ηνd)νKν(2ηνd), where η measures the correlation decay with the distance and ν is a smoothness parameter, Γ(·) is the conventional Gamma function and Kν(·) is the modified Bessel function of the second kind of order ν (see, e.g. Abramowitz and Stegun (1965)). This spatial correlation model is rather general, special cases including the exponential model

ρ(d)=exp(d) (2)

when the smoothness parameter ν = 0.5 and the ‘decay parameter’ η = 1, and the Gaussian correlation model

ρ(d)=exp(d2) (3)

corresponding to ν → ∞ and η = 1 (see, e.g., Waller and Gotway (2004), p. 279). Our theoretical development in the ensuing sections focuses on these two widely used cases of the Matèrn family.

The conditional auto-regressive (CAR) structure is also a popular choice. It has appealing theoretical properties, computational advantages and attractive interpretation (Cressie (1993)). A common CAR structure takes the form (Yasui and Lele (1997))

V=θ(IγMQ)1M=θ(M1γQ)1, (4)

where Q = {qij} is an n × n symmetric matrix; M is an n × n diagonal matrix with diagonal elements 1/qi+, with qi+ = Σj qij, −1 < γ < 1 is the spatial dependence parameter that controls the amount of information in an area provided by its neighbors, and θ is a scale parameter. The quantity qij controls the strength of connection between areas i and j, and often takes value 0 when areas i, j are not neighbors. When area i and area j are neighbors, a common choice is qij = 1 to reflect equal weights from neighbor areas. Note the flexibility of the CAR structure that allows a more general neighborhood concept than geographical proximity.

In the presence of measurement error we cannot observe X directly, but see instead its error-contaminated version W. The spatial linear mixed measurement error model is completed by assuming an additive measurement error model to relate W and X as

Wi=Xi+Ui, (5)

where Ui is the measurement error and is N(0, σu2), independent of the unobserved covariate Xi. Note that the measurement error variance σu2 often needs to be estimated using replicates or a validation data set.

Since the covariate X is unobserved, we use the structural modeling approach in the measurement error literature (Carroll, Ruppert and Stefanski (1995)) by assuming a parametric model for X and proceed with maximum likelihood estimation. The classical measurement error model often assumes X to be an independent and identically distributed Gaussian random variable. However since we are dealing with spatial data, it is likely that spatial correlation exists not only in the outcome variable Y, but also in the covariate X. Hence we assume a spatial linear mixed model for the unobserved covariate X,

Xi=α0+ZiTαz+ai+ei (6)

where the random effect vector (a1,…, an) ~ N{0,Σ(ζ)}, Σ(ζ) models the spatial correlation among the Xi, and the residuals ei are independent N(0, σe2). We assume the ai and the ei are independent of the Zi. Let W = (W1, W2, …, Wn)T, with X, Y, Z, a, b defined similarly. Note that we allow the spatial correlation structure Σ(ζ) among the Xi to be different from the spatial correlation structure V(θ) among the Yi. In practice, since X and Y both come from the same area, it is often reasonable to assume that they share the same spatial correlation structure with possibly different parameter θ and ζ.

It follows that the likelihood of the observed data Y, W conditional on Z is

L(Y,WZ)=L(YX,Z)L(WX,Z)L(XZ)dX.

Since all the conditional distributions inside the integral are Gaussian, the joint distribution of (Y, W|Z) has the closed form,

(Y,WZ)=(2n)2ln(2π)12lnΛ12(YμyWμw)TΛ1(YμyWμw) (7)

where μy = (β0 + βxα0)1 + Z(βxαz + βz), μw = α01 + Zαz and

Λ=cov(Y,WZ)=(βx2(ζ)+V(θ)+(βx2σe2+σε2)Iβx{(ζ)+σe2I}βx{(ζ)+σe2I}(ζ)+(σe2+σU2)I),

with I an n-dimensional identity matrix.

3 The Asymptotic Bias Analysis

It is of substantial interest to investigate the effect of measurement error by investigating the bias caused by ignoring measurement error, i.e., simply replacing X in model (1) by its error-prone version W. This problem, albeit common in spatial data and cautioned by many authors, is never formally addressed. Specifically, the direction and magnitude of biases in naive estimators obtained by ignoring measurement error are not well understood. The goal of this section is to study their asymptotic biases. Our asymptotic bias analysis shows that ignoring measurement error results in an attenuated regression coefficient estimator and an inflated variance component estimator.

We assume the spatial linear mixed measurement error model (1) only contains a single covariate X (no Z) with

Yi=β0+Xiβx+bi+εiXi=α0+ai+ei, (8)

where the distributions of bi, εi, ai, ei are the same as those in (1) and (6). The naive estimators of (β0, βx, θ, σε2) are obtained by simply replacing Xi with the error-prone observation Wi and fitting

Yi=β0,naive+Wiβx,naive+bi,naive+εi,naive (9)

where bi ~ N{0, V(θnaive)} and εiN(0,σε,naive2). Let Inline graphic = (1, W), βnaive = (β0,naive, βx,naive)T, Λnaive(ϑ)=V(θnaive)+σε2In, and ϑnaive=(θnaive,σε,naive2)T=def(ϑ1,ϑ2)T. The naive estimates would be obtained by maximizing the likelihood that ignores measurement error,

n2log(2π)12logΛnaive12(YWβnaive)TΛnaive1(YWβnaive). (10)

Specifically, they solve

1nWTΛnaive1(YWβnaive)=012n[(YWβnaive)TΛnaiveϑjΛnaive1Λnaiveϑj(YWβnaive)tr(Λnaive1Λnaiveϑj)]=0. (11)

We seek the probability limits of the naive estimates as functions of the true values as n → ∞; with a slight abuse of notation, these are βnaive and ϑnaive.

We resort to the increasing domain asymptotics framework when studying bias, as opposed to infill asymptotics. Zhang and Zimmerman (2005) compared these two frameworks and found that, for certain consistently estimable parameters of exponential covariograms, approximations corresponding to the two frameworks perform about equally well, but for those parameters that cannot be estimated consistently or are highly correlated, infill asymptotic approximation may be preferable. It is usually difficult to derive infill asymptotic properties, so the increasing domain asymptotic framework is used in this work.

Consider Λ in (10), which depends on ϑ=(θ,σε)=def(ϑ1,ϑ2). Let Λi = /∂ϑiΛ(ϑ) and Λij = 2/∂ϑi∂ϑjΛ(ϑ), where the differentiation is element-wise for i, j = 1,2. Now let λ1 ≤ … ≤ λn be the eigenvalues of Λ, and let those of Λi and Λij be λki and λkij for k = 1,…,n, respectively, with λ1iλni and λ1ijλnij for i, j = 1, 2. We consider the following modified regularity conditions of Mardia and Marshall (1984).

  • (c.1) lim sup λn < ∞, limsupλni<,limsupλnij<, for all i, j = 1, 2.

  • (c.2) Λi2=O(n12δ) for some δ > 0 for i = 1, 2, A=tr(ATA).

  • (c.3) A = (aij)2×2 is invertible, where for all i, j = 1,2, aij = {tij/(tiitjj)1/2} exists and tij = tr(Λ−1ΛiΛ−1Λj).

  • (c.4) lim(Inline graphic TInline graphic )−1 = 0.

These conditions ensure the growth and convergence of the information matrix from (10), which allows the usage of the general results of Sweeting (1980) to guarantee the convergence of the naive estimates. In practice, (c.1) and (c.2) are difficult to verify. However, using some basic matrix norm properties, we show in Appendix A.0 that the common geostatistical models, for example the exponential, Gaussian, and CAR models, satisfy (c.1) and (c.2). Condition (c.3) is an identifiability condition, ensuring that the variance components (ϑ1, ϑ2) are not linear dependent, which is satisfied in our settings. Condition (c.4) ensures that the observed covariates are not trivial and is satisfied for the measurement error models (5) and (6). Then if (c.1)–(c.4) hold, required limits exist (Sweeting (1980)) and satisfy the asymptotic equations (Harville (1977)),

limn1nE{WTΛnaive1(YWβnaive)}=0,limn12n[E{(YWβnaive)TΛnaiveϑjΛnaive1Λnaiveϑj(YWβnaive)}tr(Λnaive1Λnaiveϑj)]=0, (12)

where the expectations are taken under the true law of (Y, W) in (1) (omitting Z). In particular, we can calculate the asymptotic biases in the naive regression coefficients βnaive. The result is summarized in Theorem 1 and the proof is given in Appendix A.1 (on-line supplement), which can be found on-line at http://www.stat.sinica.edu.tw/statistica.

Theorem 1

(Asymptotic Biases in the Regression Coefficients) Under (c.1)–(c.4), the following hold, (i) The probability limit of the naive estimator βnaive is

βx,naive=λβx,β0,naive=β0+α0(1λ)βx, (13)

where

λ=limntr[{V(θnaive)+σε,naive2I}1{(ζ)+σe2I}]tr[{V(θnaive)+σε,naive2I}1{(ζ)+σe2I)}]+σu2tr[{V(θnaive)+σε,naive2I}1], (14)

and hence 0 ≤ λ* ≤ 1.

(ii) If Y and X have the same spatial covariance structure with different scale parameters,

V(θ)=θR,(ζ)=σ2R, (15)

where R is a known matrix, then λ* in (14) is

λ=limnl=1n(δlσe2+σ2)/(δlσε,naive2+θnaive)l=1n{δl(σe2+σu2)+σ2}/(δlσε,naive2+θnaive), (16)

where {δl} are the eigenvalues of R−1.

(iii) For regular (square) grid data and the conditional auto-regressive spatial correlation structure (4) defined using the adjacent neighborhood spatial correlation structure of Breslow and Clayton (1993),

λσ2+σe2(4+4γ)σ2+(σe2+σu2)(4+4γ); (17)

for regular grid data and an exponential or Gaussian spatial correlation structure,

0λ1σu2σu2+σe2+4σ2/(1e1/2)2. (18)

The results in Theorem 1 show that ignoring the measurement error causes the regression coefficient estimates to be attenuated. Calculations of the attenuation factor λ* can be quite complicated in general. Therefore the results in (ii) are particularly useful for numerically computing λ*, since it avoids the inversion of large matrices. Note that the eigenvalues therein do not depend on data if the spatial dependence parameter γ is known. For grid data, exponential and Gaussian correlation structures are often used. In these cases, (iii) provides a bound of the attenuation factor that can be easily computed.

We state in Theorem 2 the asymptotic bias in the naive variance component ϑnaive; the proof is given in Appendix A.2 (on-line supplement).

Theorem 2

(Asymptotic Biases in Variance Components) Suppose Y and X have the same spatial covariance structure with different scale parameters as at (15). Under (c.1)–(c.4), the asymptotic limits of the naive estimators of the spatial variance component and the residual variance satisfy,

θnaive=θ+(1λ)2σ2βx2,σε,naive2=σε2+{(1λ)2σe2+λ2σU2}βx2, (19)

where λ* is defined in (16).

Theorem 2 shows that when Y and X possess the same spatial covariance structure, a reasonable assumption in practice since they come from the same spatial area, the naive estimators of the spatial variance component and the residual variance both overestimate the corresponding true values. For more general cases when the spatial covariance structure of Y and X differ, the asymptotic limits of the naive estimators are difficult to calculate, and no analytic expressions are available.

The asymptotic relative biases in the naive estimators of the regression coefficients and the variance components, assuming the adjacent neighborhood spatial correlation structure is illustrated in Figure 1. Since the computation of λ* involves n → ∞, we approximate λ* with n = 1024 on a 32 × 32 lattice. The spatial dependence parameter γ in (4) are taken as γ = 0.2 and γ = 0.95. The regression coefficient is βx = .5, variance components are θ = 1 and σε2=.3. The parameters in the X models are α0 = 1.4, σ2=1.2, and σe2=.5. We iteratively calculate λ* using (16) and (19). In our experience, convergence is often achieved within 5 iterations. Then we obtain the expected naive estimates from (13) and (19). The bias curves for the naive estimates of βx and θ are plotted as a function of the measurement error variance σu2. It should be noted that the bias curves in fact correspond to the finite sample exact bias.

Figure 1.

Figure 1

Asymptotic relative biases in the naive estimates of βx and θ. The CAR spatial covariance structure with spatial dependence parameter γ = 0.2 and 0.95 was used. The true parameter values were βx =.5, θ = 1, σε2=0.5,σ2=1.2,σe2=.5. Variance of measurement error σu2 varied between 0 and 1.0. The two curves in each plot correspond to the spatial dependence parameter γ: ——— γ = 0.2; - - - γ = 0.95.

Figure 1 shows that the naive estimate of the regression coefficient βx is attenuated, while the naive estimate of the variance component θ is inflated. The biases increase with the measurement error variance σu2, but decrease with the spatial dependence parameter γ. The reason for the latter phenomenon is explained by the fact that stronger dependence implies that neighbor areas can provide more information, and hence the estimates are more resistant to the effect of measurement error.

4 Maximum Likelihood Estimation

We consider the large sample results for the maximum likelihood estimator for the spatial linear mixed measurement error models (1), (5) and (6). In particular, we show for some commonly used spatial models, the MLEs are consistent and asymptotically normal. To proceed, we assume mild regularity conditions on the parameter space and the observed covariate Z.

  • (d.1) The unknown parameters Ω in (1), (5) and (6) lie in a compact set of an Euclidean space.

  • (d.2) Let = (1, Z), where 1 is an n × 1 vector of 1’s. Assume lim n−1T = Z0 in probability, where Z0 is a positive definite matrix.

It follows that, for the common geostatistical models, such as the exponential, Gaussian and CAR models, the maximum likelihood estimator is consistent and asymptotically normal, as summarized in the following theorem. The proof is deferred to Appendix A.3 (on-line supplement).

Theorem 3

(Consistency and Asymptotic Normality of MLEs) Let Ω0 be the true unknown parameters in (1), (5) and (6) and Ω̂ be its maximum likelihood estimator. Suppose that Y and X have the exponential, Gaussian or CAR [eq. (4)] spatial covariance structure on regular grid. Then, under (d.1) and (d.2), Ω̂ is consistent and Γ1/2(Ω̂Ω0) → N(0, Ip) in distribution, where Γ1/2 is the Cholesky decomposition of Γ=defEΩ0{2/ΩΩT}, Ω0 is the truth, ℓ is as defined in (7), and Ip is the identity matrix of dimension of p, the dimension of Ω0

Theorem 3 does not require X and Y to have the same correlation structure, but, since X and Y both come from the same area, it may be reasonable to assume that they do. In such a situation we propose an EM algorithm to compute the MLEs; in particular, we assume the spatial covariance structures of the random effects b and a take the same form (15) with different scale parameters. The EM algorithm for a general spatial covariance structure is similar. The complete data are (Y, W, X, Z, b, a), where (Y, W, Z) are observed data and X, b, and a are missing data. The complete data loglikelihood is

(Y,W,X,b,aZ)=n2log(σε2)12σε2Yβ01βxXZβzb2n2log(θ)12θbTR1bn2log(σu2)12σu2WX2n2log(σe2)12σe2Xα01Zαza2n2log(σ2)12σ2aTR1a,

where || · || denotes the square norm.

Let = (1, X, Z), = (1, Z), β=(β0,βx,βzT)T, and α=(α0,αzT)T. At the (t + l)th step, let the estimator of β be β̂(t+1) and the estimator of α be α(t+1), and define the variance component estimates similarly. In the M step, we update the regression coefficients

β^(t+1)=E(XTXY,W,Z,ξ^(t))1E(XT(Yb)Y,W,Z,ξ^(t))α^(t+1)=(ZTZ)1ZTE(XaY,W,Z,ξ^(t)),

where E(·|Y, W, Z,ξ̂(t)) is the expectation conditional on the observed data (Y, W) with all parameters taking the values of the current estimates ξ̂(t) We update the variance components by

θ^(t+1)=1nE(bTR1bY,W,Z,ξ^(t)),σ^ε2(t+1)=1nE(Yβ^x(t+1)XZβ^z(t+1)b2Y,W,Z,ξ^(t)),σ^2(t+1)=1nE(aTR1aY,W,Z,ξ^(t)),σ^e2(t+1)=1nE(Xα^0(t+1)1Zαz(t+1)a2Y,W,Z,ξ^(t)).

In the E step, we compute the conditional expectations that appeared in the above equations. The closed-form expressions of these conditional expressions are derived and can be found in Appendix A.4 (on-line supplement). These steps can be easily implemented since all the quantities involved have closed form and no numerical integration is needed. Finally, the standard errors of the maximum likelihood estimates can be obtained by inverting the Fisher information matrix.

5 Simulation Study

Our simulation study aims at evaluating the finite sample performance of the naive estimates obtained by ignoring the measurement error and the maximum likelihood estimates obtained by accounting for the measurement error. We took the data to be on a regular grid. We considered the Y model (1) with a single covariate X. We assumed the adjacent neighborhood CAR spatial correlation structure (4) for both the random effects {bi} and {ai} in the Y and X models, neighbors being defined as the four adjacent areas for each location except for those on the edge. The weight qij was set to be 1 if areas i and j are neighbors and 0 otherwise. The spatial dependence parameter was γ = 0.95, mimicking what was obtained in the data example in the next section. The unobserved covariate X was generated under model (6) with mean 1.4 and variance components 1.2 and 0.3, respectively for the spatial covariance and residual error term. The observed error-contaminated version W was generated by adding Gaussian noise with variance σu2=0.50 to X. To generate the outcome variable Y, the regression coefficients were taken as (β0, βx)T = (0.0,2.0)T, and the variance components for the spatial covariance and residual error term were taken as 1.0 and 0.5. For each generated data set, we computed the naive estimates that ignored the measurement error and the maximum likelihood estimates that accounted for the measurement error using the EM algorithm. We varied the grid size to be 7(n = 7 × 7), 10(n = 10 × 10) and 20(n = 20 × 20). The averages and variances of the estimates from 500 replications are given in Table 1.

Table 1.

Results of a simulation study from 500 replications under the CAR model. A regular 20 × 20 grid design and an adjacent neighborhood covariance structure with γ = 0.95 were used. The true parameters were β0 = 0, βx = 0.5, θ=1σε2=0.3,σ2=1.2,σe2=0.5 and α0 = 1.4. The measurement error variance was σu2=0.5. Inside the brackets are estimated standard errors

Sample Size Parameter Mean of naive est model based Var. of naive est MSE Mean of MLE model based Var. of MLE MSE
49 β0 0.221 0.095 (0.161) 0.200 −0.045 0.150 (0.210) 0.210
βx 0.326 0.011 (0.012) 0.040 0.513 0.030 (0.043) 0.043
θ 1.141 0.355 (0.383) 0.401 1.041 0.283 (0.367) 0.368
σε2
0.461 0.054(0.063) 0.090 0.374 0.045 (0.056) 0.061

100 β0 0.261 0.062(0.072) 0.14 −0.042 0.079 (0.090) 0.09
βx 0.323 0.005 (0.005) 0.036 0.507 0.0157 (0.0157) 0.016
θ 1.066 0.222 (0.265) 0.269 0.957 0.209 (0.239) 0.240
σε2
0.399 0.0268 (0.033) 0.043 0.326 0.0312 (0.029) 0.030

400 β0 0.247 0.0160 (0.0154) 0.076 0.0032 0.020 (0.022) 0.022
βx 0.318 0.0012 (0.0012) 0.034 0.503 0.0035(0.0035) 0.0035
θ 1.015 0.067 (0.073) 0.073 0.989 0.062 (0.072) 0.072
σε2
0.376 0.0068 (0.0069) 0.012 0.304 0.0068 (0.0072) 0.0072

We next examined the performance of the MLE when the spatial correlation structure was specified to be the exponential model as well as the Gaussian model. The locations of subjects were sampled uniformly over region [0, n]2, where n is the number of subjects. We set n = 49,100,400 in our simulations. The results are documented in Tables 2 and 3.

Table 2.

Results of a simulation study from 500 replications under the Gaussian model. The locations of subjects were sampled uniformly over the region [0, n]2, where n is the number of subjects. The true parameters were β0 = 0, βx = 0.5, θ=1σε2=0.3,σ2=1.2,σe2=0.5 and α0 = 1.4. The measurement error variance was σu2=0.5. Inside the brackets are estimated standard errors

Sample Size Parameter Mean of naive est model based Var. of naive est MSE Mean of MLE model based Var. of MLE MSE
49 β0 0.276 0.107 (0.110) 0.186 −0.038 0.136 (0.191) 0.192
βx 0.320 0.010 (0.012) 0.044 0.532 0.031 (0.038) 0.039
θ 1.076 0.168 (0.180) 0.185 0.970 0.150 (0.156) 0.156
σε2
0.365 0.019 (0.018) 0.022 0.300 0.020 (0.021) 0.021

100 β0 0.218 0.043 (0.041) 0.119 −0.022 0.055 (0.057) 0.057
βx 0.338 0.0057 (0.0056) 0.032 0.512 0.0128 (0.0134) 0.013
θ 1.049 0.077(0.078) 0.081 0.982 0.071 (0.066) 0.066
σε2
0.373 0.0153 (0.0158) 0.0211 0.298 0.0150 (0.0150) 0.0150

400 β0 0.176 0.0076 (0.0084) 0.039 0.034 0.009 (0.006) 0.006
βx 0.369 0.0015 (0.0018) 0.0189 0.498 0.002 (0.002) 0.002
θ 1.027 0.026 (0.022) 0.022 1.019 0.023 (0.024) 0.024
σε2
0.386 0.013 (0.014) 0.021 0.303 0.0108 (0.0108) 0.011

Table 3.

Results of a simulation study from 500 replications under the exponential model. The locations of subjects were sampled uniformly over the region [0, n]2, where n is the number of subjects. The true parameters were β0 = 0, βx = 0.5, θ=1σε2=0.3,σ2=1.2,σe2=0.5 and α0 = 1.4. The measurement error variance σu2=0.5. Inside the brackets are estimated standard errors

Sample Size Para meter Mean of naive est model based Var. of naive est MSE Mean of MLE model based Var. of MLE MSE
49 β0 0.223 0.145 (0.179) 0.228 −0.030 0.187 (0.239) 0.239
βx 0.339 0.010 (0.012) 0.038 0.528 0.0270 (0.0323) 0.033
θ 1.079 0.220 (0.268) 0.274 0.966 0.191 (0.235) 0.236
σε2
0.408 0.0535 (0.0670) 0.079 0.364 0.052 (0.063) 0.067

100 β0 0.216 0.062 (0.073) 0.108 −0.021 0.071 (0.079) 0.079
βx 0.350 0.0057 (0.0058) 0.028 0.509 0.012 (0.013) 0.013
θ 1.037 0.114 (0.130) 0.131 0.982 0.113 (0.134) 0.134
σε2
0.399 0.037 (0.042) 0.038 0.314 0.036 (0.037) 0.037

400 β0 0.168 0.009 (0.010) 0.038 −0.002 0.011 (0.011) 0.011
βx 0.374 0.0015 (0.0015) 0.017 0.495 0.0026 (0.0026) 0.0026
θ 1.037 0.040 (0.030) 0.031 1.010 0.037 (0.045) 0.045
σε2
0.387 0.022 (0.014) 0.021 0.298 0.020 (0.024) 0.024

All the results (Tables 13) show that the naive estimate of βx is attenuated while the naive estimates for θ and σε2 are inflated, agreeing with our asymptotic bias analysis. The maximum likelihood estimates computed using the EM algorithm, on the other hand, performed very well. The mean of the estimates of the regression coefficients and the variance components were very close to the corresponding true values. As expected, there was a bias-variance tradeoff. The MLEs effectively eliminated the biases in the naive estimators but had larger variances. As an overall measure of performance using the MSE, the MLEs had smaller MSEs than the naive estimators. The MSE gain was more apparent as n increased.

Finally, to compare the empirical results with our theoretical asymptotic bias analysis results, we computed the theoretical values of the naive estimate using the results in Theorems 1 and 2. For example, under the CAR model with γ = 0.95, these values were 0.254, 0.318, 1.039, 0.367 for β0, βx, θ and σε2, compared with 0.247, 0.318, 1.027, 0.376 of the average naive estimates based on 500 simulations for grid size 20 (n=400) (see Table 1). Hence, our theoretical values do match with our simulation results.

6 Analysis of Scottish Lip Cancer Incidence Data

The Scottish lip cancer incidence data were collected in each of the 56 counties of Scotland (Breslow and Clayton (1993)). For each county, the number of lip cancer cases among males from 1975–1980 and the percentage of AFF employment in all employed population were reported. Earlier analysis found that the rates were higher in counties with higher proportion of the population employed in agriculture, forestry, and fishing (AFF) – the professions that require working outdoors. This observation reflects the biological plausible causal relationship between ultraviolet rays and lip cancer. Breslow and Clayton (1993) applied spatial mixed models to study the association between the percentage of the AFF employment and the lip cancer incidence. However the exposure of main interest is the exposure to sunlight, a known risk factor for lip cancer. The AFF employment variable serves as a surrogate for the degree of exposure to sunlight. Since we mainly focused on the association between lip cancer and the exposure to sunlight, we need to account for the measurement error in using the AFF employment variable to measure the degree of exposure to sunlight.

Breslow and Clayton (1993) modeled the standardized morbidity ratios calculated by dividing the observed number of cancer cases by the age-adjusted expected cancer cases using a Poisson regression model. To apply our methodology, we first took a square root transformation of the observed SMR; the transformed SMR approximated a normal distribution well, which was verified using the Shapiro-Wilks test. We applied the spatial linear mixed measurement error model to account for the measurement error.

Following Breslow and Clayton (1993), we assumed the adjacent neighborhood spatial correlation structure for the square-root transformed SMR. These authors also noted that the covariate, the percentage of the AFF employment, exhibited the same spatial aggregation as the SMR. We hence assumed the same spatial correlation structure with a different scale parameter for the AFF variable.

The analysis results are given in Table 4. The naive analysis showed a strong effect of the AFF employment on the SMR (βx = 0.139 and SE=0.091), and the spatial correlation seemed to dominate in the total variation (θ̂ = 0.310, σ^ε2=0.0389). We next considered the spatial linear mixed measurement error model to account for the measurement error in the AFF employment. Since no validation data set was available, the measurement error variance σu2 could not be estimated directly from the data. We fit a linear random intercept CAR model on W. This allowed us to estimate the sum of σe2 and σu2 as 0.041. We then did sensitivity analysis by varying σu2 from 0.0, naive analysis, to moderate measurement error, σu2=0.020, to severe measurement error, σu2=0.035. The estimates of the dependence parameter γ were 0.922 when σu2=0, 0.928 when σu2=0.02 and 0.932 when σu2=0.035, all of which were close to the estimate of 0.93 obtained by Yasui and LeLe (1997), and indicated a strong spatial dependence. Second, all the analyses indicated that working outdoors was associated with the risk of lip cancer. Third, ignoring measurement error did attenuate the regression coefficient estimates. As σu2 increased, the estimates of the regression coefficients became larger. For example, the estimate of the coefficient of ‘AFT’, with estimated standard error in brackets, increased from 0.132 (0.093) when σu2=0, to 0.153 (0.099) when σu2=0.02, and to 0.172 (0.104) when σu2=0.035, while the variance component for the spatial correlation part was estimated as 0.434 (0.245) when σu2=0, 0.414(0.234) when σu2=0.02, and 0.394 (0.228) when σu2=0.035. These results indicated that that accounting for measurement error increased the magnitude of the estimated effects of ‘AFT’ while it decreased the overestimation of the spatial variance component.

Table 4.

Sensitivity analysis of Scottish Lip Cancer Incidence Data: Outcome variable is the square root of SMR; the covariate is AFF/10. The measurement error variance varied between 0 (naive), 0.02 (moderate) and 0.035 (severe)

Estimate ± standard error
Parameter naive moderate severe
σu2=0.0
σu2=0.02
σu2=0.035
γ 0.922 ± 0.072 0.928 ± 0.044 0.932 ± 0.043
β0 0.939 ± 0.164 0.923 ± 0.168 0.908 ± 0.171
βx 0.132 ± 0.093 0.153 ± 0.099 0.172 ± 0.104
θ 0.434 ± 0.245 0.414 ± 0.234 0.394 ± 0.228
σε2
0.017 ± 0.045 0.021 ± 0.044 0.024 ± 0.044
σ2
1.258 ± 0.258 1.183 ± 0.256
σe2
0.0001 ± 0.0005 0.0001 ± 0.0003
α0 0.8033 ± 0.258 0.8030 ± 0.258

7 Discussion

In this paper we have proposed spatial linear mixed measurement error models to account for covariate measurement error and spatial correlation in spatial data. Our asymptotic bias analysis shows that, by ignoring the measurement error, the naive estimators of the regression coefficients are attenuated and the naive estimators of the variance components are inflated. We give formulae for calculating these biases for a general case, and provide simplified forms or bounds for some commonly-used spatial correlation structures. Our numerical calculation also shows that the biases are related to the spatial dependence parameter γ for an adjacent neighborhood structure.

We have developed a structural modeling approach to accounting for the covariate measurement error in spatial data, where spatial linear mixed models are assumed for both the outcome and the unobserved covariate, and an additive model is assumed for the observed error-prone covariate. An EM algorithm is developed to compute the maximum likelihood estimate. Our simulation study shows that the maximum likelihood estimator works well in finite samples and appropriately corrects for the bias in the naive estimator. We find that the maximum likelihood estimators correct the biases in naive estimators, but are associated with larger variances.

On the computational side, our algorithm requires operations of matrices of large size, including inversion of large matrices. We alleviate the computational burden by diagonalizing the matrices simultaneously. Since the sizes of the matrices involved increase rapidly with the grid size of the spatial areas, many operations on these matrices are needed in each EM iteration. These cause problems in handling large data sets with the EM algorithm. Here it might be more convenient to adopt an MCMC algorithm, especially if one uses the conditional autoregressive spatial covariance structure.

Ou structural modeling approach, where a parametric model is assumed for the unobserved covariate X, might be sensitive to misspecification of the distribution of X. An alternative estimation in the measurement error literature is functional modeling, such as SIMEX (Carroll, et al. (1995)), which makes no distributional assumption on X. However it can be less efficient than the MLE when the distribution of X is correctly specified. It is of interest in future research to compare these two approaches in terms of their robustness and efficiency.

We have concentrated in this paper on Gaussian spatial outcomes. Work is underway to extend the results to non-Gaussian spatial outcomes, with measurement error in the covariate, within the framework of spatial generalized linear mixed models (e.g. Diggle et al. (1998)).

Acknowledgments

The authors wish to thank the Editor, an Associate Editor and two anonymous referees for their insightful suggestions, which significantly improved this work.

References

  1. Bernardinelli L, Pascutto C, Best NG, Gilks WR. Disease mapping with errors in covariates. Statistics in Medicine. 1997;16:741–752. doi: 10.1002/(sici)1097-0258(19970415)16:7<741::aid-sim501>3.0.co;2-1. [DOI] [PubMed] [Google Scholar]
  2. Breslow NE, Clayton D. Approximate inference in generalized linear mixed models. Journal of the American Statistical Association. 1993;88:9–25. [Google Scholar]
  3. Carroll RJ, Chen R, George EI, Li TH, Newton HJ, Schmiediche H, Wang N. Ozone exposure and population density in Harris county, Texas. Journal of the American Statistical Association. 1997;92:392–415. [Google Scholar]
  4. Carroll R, Ruppert D, Stefanski LA. Measurement Error in Nonlinear Models. Champman and Hall; New York: 1995. [Google Scholar]
  5. Cressie NA. Statistics for Spatial Data. John Wiley & Sons; New York: 1993. [Google Scholar]
  6. Diggle P, Moyeed R, Tawn J. Model-based Geostatistics. Applied Statistics. 1998;47:299–350. [Google Scholar]
  7. Guyon X. Random Fields on a Network: Modeling, Statistics, and Applications. Springer-Verlag Inc; Berlin: 1995. [Google Scholar]
  8. Harville DA. Maximum likelihood approaches to variance component estimation and to related problems. Journal of the American Statistical Association. 1977;72:320–340. [Google Scholar]
  9. Lahiri S. On inconsistency of estimators based on spatial data under infill asymptotics. Sankhya. 1996;A 58:403–417. [Google Scholar]
  10. Lehman EL, Casella G. Theory of Point Estimation. 2. Springer; New York: 1998. [Google Scholar]
  11. Mardia KV, Marsh RJ. Maximum likelihood estimation of models for residual covariance in spatial regression. Biometrika. 1984;71:135–146. [Google Scholar]
  12. Prentice R, Sheppard L. Aggregate data studies of disease risk factors. Biometrika. 1995;82:113–125. [Google Scholar]
  13. Schall R. Estimation in generalized linear models with random effects . Biometrika. 1991;78:719–727. [Google Scholar]
  14. Schervish M. Theory of Statistics. Springer; New York: 1995 . [Google Scholar]
  15. Stein ML. Statistical Interpolation of Spatial Data: Some Theory for Kriging. Springer; New York: 1999 . [Google Scholar]
  16. Waller L, Carlin BP, Xia H, Gelfand AE. Hierarchical spatio-temporal mapping of disease rates. Journal of the American Statistical Association. 1997;92:607–617. [Google Scholar]
  17. Waller L, Gotway C. Applied Spatial Statistics for Public Health Data. Wiley; New York: 2004. [Google Scholar]
  18. Wang N, Davidian M. Effects of covariate measurement error on nonlinear mixed effects models. Biometrika. 1996;83:801–812. [Google Scholar]
  19. Wang N, Lin X, Gutierrez RG, Carroll RJ. Bias analysis and SIMEX approach in generalized linear mixed measurement error models. Journal of the American Statistical Association. 1998;93:249–261. [Google Scholar]
  20. Wilkinson JH. The Algebraic Eigenvalue Problem. Oxford University; Oxford: 1965 . [Google Scholar]
  21. Xia H, Carlin BP. Spatio-temporal models with errors in covariates: mapping Ohio lung cancer mortality. Statistics in Medicine. 1998;17:2025–2043. doi: 10.1002/(sici)1097-0258(19980930)17:18<2025::aid-sim865>3.0.co;2-m. [DOI] [PubMed] [Google Scholar]
  22. Yasui Y, Lele S. A regression methods for spatial disease rates: an estimating function approach approach. Journal of the American Statistical Association. 1997;92:21–32. [Google Scholar]
  23. Zhang H, Zimmerman D. Towards reconciling two asymptotic frameworks in spatial statistics. Biometrika. 2005;92:921–936. [Google Scholar]

RESOURCES